Sie sind auf Seite 1von 9

J. Chil. Chem. Soc.

, 53, Nº 2 (2008)

INTERFACE BETWEEN TWO IMMISCIBLE LIQUID ELECTROLYTES: A REVIEW

PETR VANÝSEK*1 AND LUIS BASÁEZ RAMÍREZ2


1
Northern Illinois University, Department of Chemistry and Biochemistry, DeKalb, IL 60115, USA.
2
Departamento de Química Analítica e Inorgánica, Facultad de Ciencias Químicas, Universidad de Concepción, Chile.
(Received: 13 December 2007 - Accepted: 22 April 2008)

ABSTRACT

An overview on electrochemistry on the interfaces between two immiscible electrolyte solutions is given.

Keywords: Review, ITIES, liquid-liquid interface, ion transfer, electron transfer, phase transfer catalysis, analytical applications, interfacial structure,
DIGISIM, modeling, X-ray reflectivity, double layer, immiscible liquids, electrochemistry.

BACKGROUND In order to observe ion transport across the interface due to applied
potential, it is first necessary to be able to apply appropriate potential on the
The concept of immiscibility of certain liquids, such as oil and water, must interface, i.e., the interface has to be made a polarizable interface. In order
have been known to the humankind for millennia, and it must have captivated to make a polarizable interface, it is first necessary to make the two phases
early scientists just as much as it fascinates by its specific properties scientists conductive to be able to apply potential from external electrodes. This is done
today. In our review we will focus on a specialized interface between two by dissolving suitable supporting electrolyte salts in each phase. To make the
immiscible liquids, an interface arising between two immiscible electrolyte interface polarizable, at least in a certain potential window, the salts dissolved
solutions. An electrolyte, a medium with ionic conductivity and mobile in the respective phase must be preferentially soluble in one phase, but not
charge carriers, introduces to the system of two immiscible phases additional in the other. The two salts often used are hydrophilic LiCl, which is used as
property not observed on an oil-water interface. Since the two phases are now the supporting electrolyte for the aqueous phase and tetrabutylammonium
conductive and charge between the two phases can achieve equilibrium through tetraphenylborate (TBATPB), which is lipohilic and is used as the supporting
ion transport at the interface, the ionic conductivity of both the phases imparts electrolyte for the organic phase. Nitrobenzene or 1,2-dicholoroethane (DCE)
electrical potential difference on such interface. are used very often as the organic phase. Fig. 1 shows the potential window in
which such interface is polarizable.
This interface, the interface between two immiscible electrolytes (ITIES),
has some functional similarities with other types of interfaces. An interface is
defined as a boundary between two distinct phases. Hence, there is an interface
between a liquid and the glass wall of the container holding the liquid; there is
an interface between metal and liquid (for example copper metal and aqueous
solution of copper sulfate); there is an interface between water and some
organic solvents (for example nitrobenzene or 1,2-dichloroethane).

From the electrochemical thermodynamic point of view the copper


metal/copper sulfate solution interface is treated in a fairly simple textbook
manner. The interface is described as a half cell and its potential is given by
the Nernst equation, utilizing the activity of the copper ions in the solution and
the standard reduction potential of Cu2+. When the potential of this interface
is studied in more detail it becomes obvious that the interfacial potential is
actually maintained by equilibrium between Cu2+ ions in the aqueous solution
and the activity, albeit abstract, of the copper ions in the metal. The general
reversible equilibrium

Cu2+ + 2 e Cu (1)

is the prerequisite for maintaining a steady and defined potential. In fact,


similar equilibria, e.g., between the activity of H3O+ in solution and activity of
OH- groups on surface of hydrated glass of a glass electrode is the principle of
the pH electrode. In short, any system that has an interface on which equilibrium
of charged species is dynamically established, will be a site of an interfacial
potential. By following this idea, it can be similarly envisioned that an interface
between two immiscible solutions, of which each contains certain activity of a
particular ion (K+, for example) will be a site of a potential difference driven by Figure 1. Curve of ITIES for a supporting electrolyte
the relative activities of potassium ion in each of the two phases. An example
of this is, in fact, an ion selective electrode based on a liquid membrane in The figure is a cyclic voltammogram, which shows the current flowing
which the analyte is an aqueous solution containing dissolved potassium ions; through the interface in response to the applied potential. Within the potential
the sensing side is an organic solvent. An organic solvent will not typically window, only little current flows, due mostly to charging of the interface
dissolve potassium ions. However, making them more lipophillic, for example, (charging current). Outside the window the ions of the supporting electrode
by complexation with valinomycin, can make them soluble. This is known as begin to transport into the opposite phase, contributing to the increasing
facilitated transport. background current. Since it is traditional to assign the polarity of the interface
to the aqueous phase (as if the nonaqueous phase were grounded), the right
With proper setup, the interface between the two liquids can be made into hand side of the curve in Fig. 1 corresponds to the aqueous phase becoming
a boundary that responds, for electroanalytical purposes, like an electrode. increasing more positive. As the polarization continues (1 on the curve), TPB-
However, the distinction is that instead of reduction or oxidation of a species begins to transport from nitrobenzene to water and Li+ begins to be transported
on the electrode surface, a transport of a charged species thought the interface, form water to nitrobenzene. The relative contribution of the lipophilic anion
governs the observed current flow. and the hydrophilic cation depend on the ranking of these ions on the scale
of the Gibbs energies of transfer1; since they are similar, both ions contribute

e-mail: pvanysek@niu.edu 1455


J. Chil. Chem. Soc., 53, Nº 2 (2008)

to the observed background current. After switching the direction of the scan described in a skeletal form by Fermin et al7. The process involves metathesis
(2 on the curve), tetraphenylborate (TPB-) previously transported into water of stoichiometric amounts BTPPACl and LiTPFB dissolved in 2:1 mixture
crosses again back into nitrobenzene and Li+ from nitrobenzene crosses back of methanol and water, followed by recrystallization from hot acetone. The
to water. As the cycle continues through the window of polarizability, the initial precipitation requires additional amounts of the methanol-water mixture,
current is mostly due to the charging of the interface; eventually, the aqueous therefore it is not necessary to dissolve the starting material in least amount
phase is becoming less and less positive with respect to the organic phase (3 of solvent. The solubility of the product is much higher in hot acetone than in
on the curve) and transport of Cl- from water to nitrobenzene and transport of cold, so recrystallization is pretty simple. The melting point of our product was
tetrabutylammonium (TBA+) from nitrobenzene to water is observed. Finally, 223-225 oC. It should be noted that the acetone precipitate should be washed by
after the potential switch (4 on the curve) the previously transported ions copious amounts of 2:1 mixture of methanol and water, to rinse out any starting
return to their phase of origin, Cl- crosses back to water and TBA+ crosses to material, which otherwise causes large background current.
nitrobenzene.
In the presence of an ion that can partition between the two phases it is
The potential at which the ion transfer across the interface happens is possible to obtain a voltammogram similar to that of a redox couple. Fig. 3
related to the Gibbs energy of transfer. Thermodynamically, this energy and shows such case in which a cation with potential of transfer equal to 159 mV,
the corresponding potential are normalized to the standard Gibbs energy of corresponding to the cesium ion, is being transferred across the interface.
transfer. It is a function of the particular ion, as well as a function of the solvent
pair studied. These values are available in various tables including a good web
based data base2 and Table I gives an example of some values for individual
ions transferring from water to nitrobenzene.

Table I.- Standard potentials of transfer Δα β φo for selected individual


ions from water (phase α) to nitrobenzene (phase β)

Ion Δα β φo (mV)
Li+ 395
UO22+ 373
Mg2+ 361
Na+ 354
Fe3+ 319
K+ 242
Rb+ 201
Cs+ 159
Choline+ 117
Acetylcholine+ 49
TMeA+ 35
TBA+ - 248
TPB- 372
Picrate- 47
Dodecylsulfate- - 43
ClO4- - 83
Cl- - 324 Figure 3. Transfer of a Cs+ ion - DIGISIM simulated result.

The values for individual ions cannot be fundamentally determined It should be noted that the diffusion controlled process is described by the
separately, one ion is always related to another in a chain of measurements. same equations as a redox process – transfer across the interface is considered
Therefore, an assumption has been made that in the process of determining the fast compared to the diffusion control in both phases towards and away from
values for the cation and the ion of tetraphenylarsonium tetraphenylborate both the interface, therefore the same math applies. However, the interface does not
will have equal values, based on their similar sizes 3-6. experience a redox process. The difference between the diffusion controlled
redox and ITIES situation are compared in Fig. 4.
It is desirable to have the operating window as wide as possible. Although
TBATPB is traditionally used, more lipophilic salt has recently seen increase
in use (Fig. 2), the very lipophilic

Figure 2. Bis(triphenylphosporanylidene)ammonium tetrakis (pentafluo-


rophenyl)borate

bis (triphenylphosporanylidene) ammonium tetrakis (pentafluorophenyl) borate


(BTPPATPFB). This salt has desirable wider potential window than TBATPB.
Its disadvantage is that it is rather expensive and so far it has to be prepared
rather than purchased. The preparation is not particularly involved and is Figure 4. Comparison between ITIES and a redox system

1456
J. Chil. Chem. Soc., 53, Nº 2 (2008)

In case A when negative potential is applied at the working electrode, applications in analytical chemistry and electrochemistry27-30.
electron transfer takes place at the solution/metal electrode interface, iron(III)
is reduced to iron(II) and an electron moves “upwards” from the electrode to the The current associated with the ion transport across the interface is
now-reduced iron ion. As a consequence, negative charge (arrow) flows through governed by the same mass transport limitations as are redox processes on
the outside electrical circuit. In case B, when negative potential is applied at the a metal electrode/solution interface, namely, in unstirred systems with
bottom (nonaqueous) phase in which ion A- is present, the negatively charged fast electron transport the current governing step is diffusion. As long
ion moves “upwards” from the nonaqueous phase to the aqueous phase. As as the transport of the ion on ITIES is fast (which it usually is) then the
a consequence, negative charge (arrow) flows through the outside electrical current associated with the ion transport is governed by the same diffusion
circuit. Although the interfacial process in case A and B are different, the same equations. Therefore, we used successfully the modeling software DIGISIM31
effect, current flow in the external electrical circuit, is observed. (Bioanalytical Systems) to generate voltammograms which well agree with the
voltammograms obtained from an experiment. It is important to realize that
The potential on the interface, as governed by a single ion that can the potential window in ITIES is much narrower than is the typical working
partition between the two phases in described by equation range of an ideally polarizable electrode. Therefore, the potential of transfer
of the supporting electrolytes has to be included in the CV properties as well.
To generate a curve of the supporting mechanism we chose four mechanisms;
(2) reduction of a species by one electron, with the redox potential being equal
to the standard potential of transport for the supporting electrolytes, i.e., Cl-,
Li+, TBA+ and TPB-. In practice, it is sufficient to include only the cation and
which is similar in its appearance to the Nernst equation and in fact, it can be anion which are the more restrictive, in this case TBA+ and TPB-. To visualize
derived by using the Nernst formalism. It is useful in situations where the other transport of a semihydrophilic ion across an interface, additional mechanism
ions, including the counterions, are well confined in their respective original (again, a reduction) is added, with the redox potential Eo in the software set
phases. However, when more that one ion participates in the equilibrium, the to the standard potential of transfer for the semihydrophilic ion. Fig. 3 shows
equation becomes rather complicated8: such simulation. The potentials for the restrictive ions of the supporting
electrolytes were set to 0.372 V (TPB-) and -0.248 V (TBA+) [Compare to the
standard potentials in Table I.]. The concentration of these ions was 0.1 mol/
(3) l, value typical for such experiments. The potential of the semihydrophobic
ion was 0.159 V, chosen for illustration to fall between the potentials of the
supporting electrolytes. This corresponds to the potential of transport of cesium
ion between water and nitrobenzene. Its concentration was 1.0 mmol/l. The
scan rate was 25 mV/s. The actual scan was between –0.143 and + 0.272 V.
where Vα and Vβ are the volumes of the phases α and β (usually water, oil), Additional adjustable parameter is the diffusion coefficient of the ion. We used
Δφ is the interfacial potential, Δφoi is the standard potential of transfer for the the default value, 1x10-5 cm2/s, which is somewhat higher than would be the
individual ion (for example listed in Table I), m is the number of moles of each actual value. However, since the diffusion coefficient enters into the equations
of the ions, n is the charge (signed) of the respective ion, and γ is the activity of as a square root, the results are not very sensitive to the exact value and for
the ion in the respective phase, α or β. For more than two species this equation demonstration purpose of suitability of DIGISIM to simulate ITIES curves this
cannot be solved explicitly. However, with the help of an iterative solver it can is adequate.
be successfully solved and used to calculate the interfacial potential from the
known values of the standard potentials, or, it can be used also to solve for a An example of the electrochemical cell used for the experiments with
particular unknown standard potential, if the partition coefficients are known9. ITIES is shown in Fig. 5. It has some degree of complexity, because the issue
of a 4-electrode potentiostat and the issue of a reference electrode have to be
When the immiscible phases are in contact for sufficient length of time, addressed.
equilibrium according to the equation (3) will be established and there will
be no net current flow through the system. However, when such interface is In typical electrode electrochemistry a 3-electrode potentiostat is used,
polarized from an external source, new equilibrium has to be established and with the working, the counter and the reference electrode. In principle, two
this can happen only trough reequilibration of the phases, by ion transport from electrodes are needed; the 3-electrode setup allows the reference electrode not
one phase to another and therefore, by current flow. In the most basic form the to pass any current and therefore avoid polarization, and the current is supplied
expression describing the current at the interface is by the counter (or auxiliary) electrode. Such setup allows to compensate for
the resistance of the solution. In ITIES we have to content with two sources
of resistance, both the aqueous and the nonaqueous solutions, with the ITIES
(4) (functional equivalent of the electrode) sandwiched between them. Therefore
a special potentiostat is used, which has input for 2 reference electrodes, rather
than for one. Additionally, two counter electrodes are used. A number of
commercial potentiostats allow this connection either directly or after some
where A is the interfacial area, F is the Faraday constant, k is the formal modifications. Solartron 1286 or 1287 is an example of the instrument used in
rate constant of the transfer, α is the charge transfer coefficient, ci(w) and ci(o) our laboratory.
are the concentration of ion i in the aqueous and the oil phase respectively,
is the standard potential of transfer of ion i and is the potential difference The cell is operated in such a way that the reference electrode inputs
applied on the interface. This equation is in its formalism the same as the from the instrument are connected to the cell reference electrodes, which then
Butler-Volmer equation written for redox processes. A more involved equation extend as Luggin capillaries to the vicinity of the L/L interface, marked ITIES
can be written to include the Frumkin correction, which, as has been shown by in the diagram. The two counter electrodes are connected to platinum flag
d’Epenoux et al10., applies also to ITIES. This more complex equation can be electrodes separated from the working solutions by a glass frit that prevents any
found for example in the review publication11. electrochemical products formed on the counter electrodes from contamination
of the interface. It should be noted that although there is no redox process
The interfaces between two immiscible electrolytes are of continuing interest occurring on the interface, as long as there is current flowing through the cell,
to many researchers, because of their relevance to such diverse applications such redox processes (usually oxidation or reduction of the solvents) is taking place
as ion-pairing12,13, charge-transfer14,15, adsorption-desorption15, complexation16, on the platinum surface of the counter electrodes.
extraction17 acid-base processes18, catalysis19, micellar chemistry20, modeling
of interactions at biological cell membranes21-24, solvation dynamics25 and The potential of the whole cell, which includes the interface, is monitored
fundamental studies of the nature of such interface26. by the pair of the reference electrodes. The aqueous reference electrode is
usually simple Ag/AgCl electrode, which works well, because in most cases
Ion distributions in electrolyte solutions near charged interfaces underlie the solution contains chloride, which pins the Ag/AgCl potential to a defined
processes as diverse as electron and ion transfer at biomembranes and redox value. The reference electrode for the non-aqueous phase is more involved.
processes at mineral-solution interfaces, and also influence many practical It is a system with two interfaces. The actual metallic connection is realized

1457
J. Chil. Chem. Soc., 53, Nº 2 (2008)

by an Ag/AgCl electrode, immersed in aqueous solution of chloride. This ion transfer in this case is quasireversible, because following the permanganate
chloride has a counterion which is the same as is the cation in the nonaqueous transport into the organic phase a chemical reaction occurs. The kinetic
phase. Therefore, if the nonaqueous solution contains TBATPB, then the parameters were obtained by cyclic voltammetry and chronopotentiometric
aqueous solution will contain TBACl. This solution is then in contact with the techniques.
nonaqueous solution forming a reference interface; the two solutions in contact
have a common cation (TBA+ in this case), which equilibrates between the two For many applications and even for theoretical calculations dealing with L/
phases and sets up the interfacial potential according to equation (2). Therefore, L interface it is necessary to know the diffusion coefficients and the transferring
the potential applied by the potentiostat and reported on the voltammograms species and also the effective charges of the transferred species. Yuan et al38.
is not usually the “standard potential of transfer;” rather, it is a potential that is demonstrated how to do this by a chronoamperometric method. To this end
the sum of the interfacial potential, the potential of the two reference electrodes they employed a micropipette electrode. Since the micropipette has a large time
and the potential of the reference interface. constant (due to high resistance and relative large capacitance of the thin glass
The interface should be positioned between the two Luggin capillaries. surrounding the pipette), only measurements at times more than 5 ms were
Different means of achieving this are possible. We are using a screw driven possible. The authors performed finite element simulation to show validity of
piston that allows fine change of volume in the lower part of the cell, facilitating the experimental data. For protamine they determined diffusion coefficient
thus adjustment of the interface. to be (1.2± 0.1) x10-6 cm2/s, with ionic charge +20±1, which is close to the
excess positive charge of the molecule. For the ETH129 calcium ionophore
they determined that each ETH molecule39 transfers +0.67 charge per each Ca2+
ion, and +1 charge per Mg2+ transfer, which corresponds to formation of 1:3
complex for calcium and 1:2 complex for magnesium, in agreement with other
measurements. They also40 introduced the ionophore dinonylnaphthalenesul
fonate to facilitate transport of protamine at the L/L (water/dichloroethane)
interface.

The facilitated transport is very useful in situation where the ion itself falls
outside the potential window of the supporting electrolytes, usually because
it is too hydrophilic. To make the ion more oil soluble, complexation with
large, usually neutral species, is performed. Besides the already mentioned
analytical applications32 many other determinations with facilitated transport
were reported, of which only selected few can be mentioned16, 41-49. Biological,
physiological and pharmaceutical applications

Antibiotics are one class of compounds that have enjoyed particular


attention of the analytical work on the L/L interfaces. The typical function
of an antibiotic involves facilitated transport of ions (even though, it is
actually the ions, that facilitate the transport of the antibiotics) across a
biological membrane. Therefore, antibiotics with appropriate modification
will be transported across the ITIES. One of the early work demonstrated
determination of monensin, useful in synthesis of this compound for cattle
feed50-54, or nigericin55,56. Valinomycin, which forms very selective complex
with potassium, was studied on such ITIES57-63 as were β-lactam antibiotics and
their derivatives64, the channel former alamecithin65. It is not without interest
that antibiotics can be also synthesized, using a two phase method, on a liquid-
liquid interface66,67

The interface is also a natural site where polar and in particular large
molecules can be absorbed. Phospoholipids68-71, phosphatidylcholine21, 72-
74
, acetylcholine75-77, cellular protein annexin 78, and proteins (bovine serum
albumin), were all studied.
Figure 5.- Diagram of the four-electrode cell used for ITIES studies
Jänchenová et al23, 24, studied adsorption and ion pairing interactions of
phospholipids on the water-1,2-dichloroethane interface. In particular, they
Analytical applications were interested in dipalmitoyl phosphatidyl choline (DPPC) (L-α-lecithin),
which appears to have rather complex behavior.
Sun and Vanýsek32 demonstrated that the interface could be used for The authors propose the following sequence of steps23 .
determination of lead (II) ion by its transport across the interface. Because A A A
lead (II) itself is quite hydrophilic, the transport must be facilitated by a ligand L (o)
L (abs)
+ H+(w) HL+(ads) HL+(o) + A-(o) 6 HLA(o) 6 HA(o) + L (o)
(5)
former, such as polyethylene glycol.
where L is the zwitterionic form of DPPC and HL+ is its protonated form.
A

The class of compounds seeing recent interest, the dendrimers, were also The five steps indicate that the process depends both on the potential difference
investigated on the liquid-liquid interfaces33. In particular, it was the non- on the interface as well as the pH value. In the absence of multivalent ions a
redox active species, poly(propylenimine) and poly(amidoamine), for which monolayer is formed. However, in the presence of cerium(IV) sulfate it was
transfer across (acidified water)/1,2-dichloroethane interface was a viable found that the DPPC forms multilayers, as indicated by slow transport. In fact,
electroanalytical technique, since redox voltammetry is not possible. ITIES the multilayers strongly slow down, even prevent, the transport of larger ions
voltammetry allowed low micromolar detection of dendrimers. It was observed such as TMeA+, PF6-, or K+ complexed with a crown-ether. Studies of such
that the electrochemistry depended on the dendrimer family, the generation biologically important compounds as dopamine79 and promazine80 were also
number, and the experimental pH. described.

ITIES can be also successfully used for liquid-liquid extraction17, 34, 35. Jain Amemia presented several analytical papers where the property of the
et al36. demonstrated the use of calixarene compound to preconcentrate and L/L interface was used to detect species of biological importance. In81 he
transport lanthanum(III) ion. Transfer of permanganate ion37 was investigated demonstrated how meparin, negatively charged polysaccharide, can be detected
across the water-nitrobenzene interface with reported potential transfer, Gibbs on the 1,2-dichloroethane interface. The detection requires use of ionophores to
energy of transfer, the transfer rate constant and the apparent α coefficient accomplish the transfer. One of the more successful ionophores was octadecyl
(symmetry factor equivalent in the redox electrochemistry) of this reaction. The trimethylammonium ion (as bromide). Unlike nonionic ionophore, this cationic

1458
J. Chil. Chem. Soc., 53, Nº 2 (2008)

ionophore function is potential-dependent. To further increase detection limit, of molecular ordering at the interface27, 28, 193. However, there are few techniques
the authors used preconcentration stripping method and achieved detection limit that are capable to probe directly this interface on the molecular length scale.
of 0.012 unit/ml, significantly less than is the therapeutic (anticoagulant) value Such techniques include surface second harmonic generation137, 194-203, total-
(> 0.2 unit/ml). In82 he used protamines as a model species to demonstrate, first internal reflection spectroscopy204-207, ac-potential modulation spectroscopy84,
time ever, voltammetric observation of phase transfer of biological polyions at 197
, vibrational sum-frequency spectroscopy208, and time-resolved quasi-elastic
water-nitrobenzene interfaces. light scattering209, as well as neutron reflectivity210-213. To answer fundamental
questions about the structure and transport across this interface on the molecular
The smooth, unrestrained liquid/liquid interface is of great advantage in the length scale a method using X-ray surface scattering was developed214-219. This
x-ray studies on ITIES and it is certainly of some advantage in electroanalytical X-ray method provides information on interfacial molecular ordering on the
applications where the surface area is that of the geometrical area. However, sub-nanometer length scale that is complementary to that provided by the
in applications where large surface area is needed, such as in phase transfer electrochemical and optical techniques.
catalysis or in the use for energy applications in possible solar cells 82-87, the
limited surface area is a problem. Girault et al88 demonstrated that increase Fundamental Studies
in surface are can be performed when a 3-dimensional ITIES experiment is
performed on vitreous carbon. Electron transfer. Most of the analytical applications deal with transport
of ions across the interface, either directly or in some facilitated form using
An interesting nanotechnology procedure using liquid-liquid interfaces a ligand to adjust the Gibbs energy of transport appropriately to fit into the
was demonstrated by Glaser et al89, where the hexane-water interface was used potential window. In fact, in the introduction we have made a clear distinction
to align forming particles in a manner that a particle sphere consisting from two between metal electrode and solution and an interface between the electrolytes,
different materials, one on each side, is formed. These particles, called Janus and we pointed out that there is no redox reaction taking place at the ITIES.
particles after the Greek god with a face both on front and back of his head, However, in addition to ion transport across the ITIES, is also possible to
were formed from simultaneous growth of Au and Fe3O4. At least theoretically observe processes involving electron transfer. Such processes typically require
Kornyshev et al90 suggested a principle of operation of a molecular device one redox system in each of the two phases and then the electron transfer can
on ITIES that could transform the energy of light into repetitive mechanical occur on the interface. The redox couples used were for example ferrocene-
motion. ferricinium for the nonaqueous phase and ferri-ferrocyanide in the aqueous
phase220. Other work investigating transport of electrons on ITIES, using
When Nernst91,92 postulated the thermodynamic basis for electrode microinterfaces and scanning electrochemical micropipette methods, were
equilibrium potential, leading now to what is known as the Nernst equation, reported127, 130,220-229.
he also carried out with Riesenfeld experiments on liquid/liquid interfaces93.
However, early work on liquid interfaces was mostly non-electrochemical, Interfacial Structure of Pure Interfaces. There are essentially two
focusing on extraction processes, salting-out in ion solvent extraction, basic, but opposing views of the structure of the interfacial structure of the
measurements of physical properties such as interfacial tension, and liquid/liquid boundary. One model predicts a molecularly sharp interface only
physiological studies on model membranes94-102. Systematic electrochemical disturbed by contribution from capillary wave fluctuations. The other model
treatment did not begin until the late 1970’s when Koryta et al103 demonstrated assumes interface in which regions of molecules originating from both phases
that the liquid/liquid interface lends itself to the same formalism as a solution/ exist. These two views result in two very different chemical environments
metal interface and that similar, if not identical, experimental methodology for molecules at the interface. Interfacial tension and electrical capacitance
could be used. This led soon to development of various electrochemical measurements at the interface between two electrolyte solutions have provided
techniques to study the liquid/liquid interface, including, among others, studies arguments both for and against the sharp interfacial region 128, 230-237.
of the solvent dropping interface103-108, and studies of cyclic voltammetry108,
impedance measurements109-112, drop pressure method113, galvanostatic pulse Either view of the interfacial structure has also appeared in the theoretical
method114, 115, stripping voltammetry114, voltfluorometry 116-119, and transport literature129, 132, 133. These studies include a density functional approach that
across a microinterface120-125. Electron transfer and photoinduced electron leads to a mixed interfacial region at the interface with a thickness of several
transfer have been also observed on ITIES or theoretically treated19, 84, 86, 87, 126- solvent diameters238 and lattice-gas calculations that incorporate a mixed
133
, as well as electrochemical catalysis19, 126, 127, 134, 135, adsorption15, 72, 136-138 and interfacial region whose thickness is proportional to the miscibility of the two
electrodeposition26,139-154. solvents237, 239. Computer simulations based on molecular dynamics predict that
the interface should be locally sharp, only roughened by capillary waves214-240-
Newer techniques have been more recently applied to ITIES, such as the 247
that originate form thermal movement of all particles in the matter.
quartz crystal microbalance155, 156, and scanning electrochemical microscopy on
liquid/liquid interfaces 75, 157-161. The sharpness of the interface is not only a matter of academic curiosity.
It has also important practical electrochemical consequences. For example,
Although the field of liquid/liquid electrochemistry is still relatively new, Marcus calculated rate constants for electron transfer between two redox species
more scientists are finding its results or methodology relevant and important in opposite phases, using either the sharp or the diffuse interface model131, 132.
to their own work27-29. Recent efforts have led to many practical applications His two results differed by two orders of magnitude. However, at that time the
in analytical chemistry and electrochemistry 162, 163. Charge transport across scarce experimental data available and their experimental uncertainty, failed to
an interface between two immiscible ionically conductive media is very provide definitive answers to the question of the interfacial structure129.
important both in naturally occurring systems and in designed applications.
Examples include ion transport across biological membranes15, 164, 165, drug The cause of disagreement on the nature of the interfacial structure was in
delivery166, behavior of ion-selective electrodes with liquid membranes and part due to a lack of techniques that can directly probe the interface. Interfacial
similar sensors167, extraction processes in oil recovery168 or nuclear waste tension and capacitance, which were both used, probe only macroscopic
reclamation and recovery169, 170, phase transfer catalysis in organic synthesis171, properties of the interface. The interfacial width of a nearly pure interface,
pharmacology172-174, and many applications in electroanalytical chemistry130, 163, DCE/water(0.05 M KOH), which was measured by neutron reflectivity, was
174-178
, applied or developmental. Fundamentally, the interface is equivalent to found to be less than or equal to 1000 pm210. A vibrational sum-frequency
one side of a membrane of an ion selective electrode and thus these studies spectroscopy was used to study the DCE/water interface and it measured a small
usually draw on the work of ITIES as well. Microinterfaces are another useful optical response from polarization normal to the interfacial plane. Although
analytical tool, one already used to build sensors179 or to be used in the scanning this result was interpreted as evidence for a mixed interfacial region, this
pipette electrochemical microscopy180-182 or in other analytical detection technique does not directly measure the interfacial width208. Schlossman et al.
schemes 121, 183-186. Ionic liquids which recently gained popularity in chemistry used X-ray reflectivity to measure the width of the 2-heptanone/water interface
research also show promise for applications in the ITIES work187-193. to be 700±20 pm215, which agrees well with the value 730 pm predicted from
capillary wave theory. This theory describes a fluctuating, molecularly sharp
Studies of charge transport, mostly ion but also electron, across the liquid/ interface. Molecular dynamics simulations are also in quantitative agreement
liquid interface are often interpreted in terms of molecular or ionic ordering with Schlossman et al. results244. Schlossman et al. also measured the width of
at the interface. Both computer simulations and analytic theories aiming to the nitrobenzene/water interface at four different temperatures 214, 216. It turns out
understand these electrochemical studies predict or assume very often existence that the measured interfacial width is actually significantly smaller than the one

1459
J. Chil. Chem. Soc., 53, Nº 2 (2008)

calculated from the capillary wave theory. For example, at 25 °C the measured About the authors:
width is 450±10 pm and the prediction is 520 pm. To explain this discrepancy,
the computer simulations suggest the presence of both weak molecular layering Luis Basáez Ramírez is Assistant Professor at the Department of
(of nitrobenzene at the interface) as well as dipole ordering parallel to the Analytical and Inorganic Chemistry, University of Concepción, Casilla 160-C,
interface244. It was shown that either layering or a bending rigidity, which could Concepción, Chile. He received his Ph.D. in 1995 from the Catholic University
result from dipole ordering, could explain these measurements214. The spatial of Valparaiso. He is presently appointed at the University of Concepción, in
resolution of the measurements did not allow to distinguish between these 1996-2001 as an Instructor and since 2001 as an Assistant Professor.
two possibilities. However, these results unequivocally demonstrate that the
interface is molecularly sharp. Petr Vanýsek is a Professor at the Chemistry and Biochemistry Department
of Northern Illinois University. His interests are in the field of liquid interfaces
If the model of a mixed solvent region at the interface would apply, the as well as sensors, impedance studies in general and in particular impedance
width of the interfacial region would have a contribution both from thermal studies on components of fuel cell. He is particularly interested in interpreting
fluctuations (capillary waves) as well as from the thickness of the mixed the impedance data correctly, with emphasis on unearthing experimental
region. These effects would lead, however, to a substantially larger width than artifacts. He is the Secretary of the Electrochemical Society and the Regional
was measured. Therefore, the results214 are not consistent with the theory of a Representative (USA) of the International Society of Electrochemistry. He
mixed solvent region. A recent measurement on water/hexane interface using received his Ph.D. in Physical Chemistry from the Czechoslovak Academy of
neutron reflectivity248, 249 confirmed X-ray results done on the same interface250. Sciences in Prague.
Although the water/hexadecane interface cannot be used for liquid/liquid
electrochemistry, the agreement of the results from both methods (X-rays, REFERENCES
600±20 pm; neutrons, 600±10 pm) confirm soundness of the reflectivity
measurements methods. 1. B. Behr, J. Gutknecht, H. Schneider, J. Stroka, J. Electroanal. Chem. 86,
289, (1978).
The interface between two liquid can have unusual electrowetting 2. H. H. Girault, http://lepa.epfl.ch/cgi/DB/InterrDB.pl. (Page verified
properties251. In fact, unusual effect of interfacial motion related to this effect active 17 April 2008).
has been described by many before252-259. 3. M. H. Abraham, T. Hill, H. C. Ling, R. A. Schultz, R. A. C. Watt, J.
Chem. Soc., Faraday Trans. I. 80, 489, (1984).
Ion Distributions at Interfaces. The liquid-liquid interface has two 4. O. Popovych, Crit. Rev. Anal. Chem. 1, 73, (1970).
important advantages over other interfaces if one wants to study ion distribution 5. J. Rais, Collect. Czech. Chem. Commun. 36, 3253, (1971).
near a surface. First is that the fluid interface does not impose an external 6. A. J. Parker, Chem. Rev. 69, 1, (1969).
structure on the adjacent liquid, unlike what would be expected from atomic 7. D. J. Fermin, H. D. Duong, Z. Ding, P.-F. Brevet, H. H. Girault, Phys.
size patterns on a solid surface. Second, if a solid surface or even a Langmuir Chem. Chem. Phys. 1, 1461, (1999).
monolayer on the water surface were to be used, their own bound charges, 8. L. Q. Hung, J. Electroanal. Chem. 115, 159, (1980).
which are not known, would have to be separately determined. 9. P. Vanysek. Determination of Gibbs energies of transfer of extremely
lipophilic supporting electrolytes in liquidliquid electrochemistry. in The
Ion distributions near a charged, planar surface can be predicted to some 58th Annual Meeting of the International Society of Electrochemistry.
degree by Gouy-Chapman theory, which solves the Poisson-Boltzmann 2007. Banff: ISE.
equation with simplifying assumptions260, 261. One limitation of this theory is that 10. B. D’Epenoux, P. Seta, G. Amblard, C. Gavach, J. Electroanal. Chem.
it assumes point-like volume-less ions, which interact through their mean field 99, 77, (1979).
in a structureless, continuum solvent. Extensive development and modification 11. P. Vanysek, Electrochemistry on Liquid/Liquid Interfaces. Lecture Notes
of the theory has addressed the limitations of the Gouy-Chapman theory262 and in Chemistry, ed. G. Berthier. Vol. 39. Springer-Verlag. Berlin. 1985, 106
predicted that for monovalent ions the deviations are largely a result of the pp.
difference between the interfacial and the bulk liquid structure. However, only 12. A. K. Kontturi, L. Kontturi, J. A. Manzanares, S. Mafe, L. Murtomaeki,
few experimental probes are directly sensitive to the structure near the charged Ber. Bunsen-Ges. Phys. Chem. 99, 1131, (1995).
interface, and the limit of validity of Gouy-Chapman theory has not been 13. R. D. Webster, D. Beaglehole, Phys. Chem. Chem. Phys. 2, 5660,
properly tested. The X-ray structural measurements, which demonstrate the (2000).
failure of the Gouy-Chapman theory, are in agreement with predictions based 14. S. Senthilkumar, R. A. W. Dryfe, R. Saraswathi, Langmuir. 23, 3455,
upon a molecular dynamics simulation that includes the effects of interfacial (2007).
liquid structure217. 15. S. Amemiya, X. Yang, T. L. Wazenegger, J. Am. Chem. Soc. 125, 11832,
(2003).
Classical electrochemical measurements at the liquid/liquid interface have 16. M. H. M. Cacote, C. M. Pereira, L. Tomaszewski, H. H. Girault, F. Silva,
revealed inadequacies of the Gouy-Chapman theory as well139, 233, 234, 237, 263-268. Electrochim. Acta. 49, 263, (2004).
A so-called Stern layer of preferentially adsorbed solvent molecules or ions 17. B. L. Rivas, S. Villegas, B. Ruf, I. M. Peric, J. Chil. Chem. Soc. 52, 1164,
is often used to explain measurements at the electrode/solution interface269. (2007).
The modified Gouy-Chapman-Stern theory includes the adsorbed layer plus 18. V. Gobry, G. Bouchard, P.-A. Carrupt, B. Testa, H. H. Girault, Helv.
the diffuse ion distribution described by the original Gouy-Chapman theory. Chim. Acta. 83, 1465, (2000).
Preferential adsorption of ions can occur at the liquid-liquid interface, although 19. R. Lahtinen, H. Jensen, D. J. Fermin, Catalysis and photocatalysis
tension measurements show that this does not occur on the nitrobenzene with at polarized molecular interfaces. An electrochemical approach
TBATPB and water with TBABr231. to catalytic processes based on two-phase systems, self-organized
microheterogeneous structures, and unsupported nanoparticles, in
CONCLUSIONS Interfacial Catalysis, A. G. Volkov, Editor. CRC Press, Boca Raton,
2003, p. 611.
It was out intention to outline the principle of the ITIES in electrochemistry 20. C. Gamboa, R. Barraza, A. F. Olea, J. Chil. Chem. Soc. 49, 303, (2004).
and to highlight some more recent or pressing aspects of this field. For better 21. T. Kakiuchi, T. Kondo, M. Kotani, M. Senda, Langmuir. 8, 169, (1992).
in-depth understanding several larger reviews are available 10, 129, 138, 271-274. 22. Y. Yoshida, K. Maeda, O. Shirai, J. Electroanal. Chem. 578, 17, (2005).
23. H. Jaenchenova, A. Lhotsky, K. Stulik, V. Marecek, J. Electroanal.
ACKNOWLEDGEMENTS Chem. 601, 101, (2007).
24. H. Jaenchenova, K. Stulik, V. Marecek, J. Electroanal. Chem. 604, 109,
PV acknowledges support from NSF-CHE0315691. (2007).
Basáez agradece al proyecto DIUC N° 207.021.025-1.0 and DIUC N° 25. G. Moakes, J. Janata, Acc. Chem. Res. 40, 720, (2007).
208.021.025-1.0 26. A. Trojanek, J. Langmaier, Z. Samec, J. Electroanal. Chem. 599, 160,
(2007).
27. A. G. Volkov, D. W. Deamer, Liquid-Liquid Interfaces: Theory and
Methods. 1996, 421 pp.

1460
J. Chil. Chem. Soc., 53, Nº 2 (2008)

28. A. G. Volkov, Editor, Liquid Interfaces in Chemical, Biological, and 75. R. Gulaboski, C. M. Pereira, M. N. D. S. Cordeiro, I. Bogeski, E.
Pharmaceutical Applications. [In: Surfactant Sci. Ser., 2001; 95]. 2001, Ferreira, D. Ribeiro, M. Chirea, A. F. Silva, J. Phys. Chem. B. 109,
853 pp. 12549, (2005).
29. H. Watarai, N. Teramae, T. Sawada, Editors, Interfacial Nanochemistry: 76. B. Yu, B. Huang, P. Li, Microchem. J. 52, 10, (1995).
Molecular Science and Engineering at Liquid-Liquid Interfaces. 2005, 77. V. Marecek, Z. Samec, Anal. Lett. 14, 1241, (1981).
321 pp. 78. E. Bitto, M. Li, A. M. Tikhonov, M. L. Schlossman, W. Cho, Biochemistry.
30. H. H. Girault, Analytical and physical electrochemistry. EPFL Press. 39, 13469, (2000).
Lausanne, Switzerland. 2004 pp. 79. A. Berduque, R. Zazpe, D. W. M. Arrigan, Anal. Chim. Acta. 611, 156,
31. A. W. Bott, S. W. Feldberg, M. Rudolph, Curr. Sep. 13, 108, (1995). (2008).
32. Z. Sun, P. Vanysek, Anal. Chim. Acta. 228, 241, (1990). 80. L. M. A. Monzon, L. M. Yudi, Electrochim. Acta. 53, 2217, (2008).
33. A. Berduque, M. D. Scanlon, C. J. Collins, D. W. M. Arrigan, Langmuir. 81. J. Guo, Y. Yuan, S. Amemiya, Anal. Chem. 77, 5711, (2005).
23, 7356, (2007). 82. S. Amemiya, X. Yang, T. L. Wazenegger, J. Am. Chem. Soc. 125, 11832,
34. A. Oliva, A. Molinari, C. Avila, M. F. Flores, J. Chil. Chem. Soc. 51, 865, (2003).
(2006). 83. D. J. Fermin, Z. Ding, H. D. Duong, P. F. Brevet, H. H. Girault, Chem.
35. C. G. Gomez, C. R. von Plessing, G. M. C. Godoy, R. H. Reinbach, R. R. Commun., 1125, (1998).
Godoy, J. Chil. Chem. Soc. 50, 479, (2005). 84. D. J. Fermin, Z. Ding, H. D. Duong, P.-F. Brevet, H. H. Girault, J. Phys.
36. V. K. Jain, S. G. Pillap, H. C. Mandal, J. Chil. Chem. Soc. 52, 1177, Chem. B. 102, 10334, (1998).
(2007). 85. D. J. Fermin, H. D. Duong, Z. Ding, P. F. Brevet, H. H. Girault,
37. H. Heli, M. Shamsipur, M. F. Mousavi, Pol. J. Chem. 80, 313, (2006). Electrochem. Commun. 1, 29, (1999).
38. Y. Yuan, L. Wang, S. Amemiya, Anal. Chem. 76, 5570, (2004). 86. D. J. Fermin, H. Dung Duong, Z. Ding, P. F. Brevet, H. H. Girault, Phys.
39. U. Schefer, D. Ammann, E. Pretsch, U. Oesch, W. Simon, Anal. Chem. Chem. Chem. Phys. 1, 1461, (1999).
58, 2282, (1986). 87. D. J. Fermin, H. D. Duong, Z. Ding, P.-F. Brevet, H. H. Girault, J. Am.
40. Y. Yuan, S. Amemiya, Anal. Chem. 76, 6877, (2004). Chem. Soc. 121, 10203, (1999).
41. Y. Shao, M. V. Mirkin, Anal. Chem. 70, 3155, (1998). 88. S. Tan, M. Hojeij, B. Su, G. Meriguet, N. Eugster, H. H. Girault, J.
42. R. Zazpe, C. Hibert, J. O’Brien, Y. H. Lanyon, D. W. M. Arrigan, Lab Electroanal. Chem. 604, 65, (2007).
Chip. 7, 1732, (2007). 89. N. Glaser, D. J. Adams, A. Boeker, G. Krausch, Langmuir. 22, 5227,
43. P. O’ Dwyer, V. J. Cunnane, J. Electroanal. Chem. 581, 16, (2005). (2006).
44. G. Pieruz, P. Grassia, R. A. W. Dryfe, Desalination. 167, 417, (2004). 90. A. A. Kornyshev, M. Kuimova, A. M. Kuznetsov, J. Ulstrup, M. Urbakh,
45. B. Kralj, R. A. W. Dryfe, J. Electroanal. Chem. 560, 127, (2003). J. Phys.: Condens. Matter. 19, 375111/1, (2007).
46. D. Zhan, Y. Xiao, Y. Yuan, Y. He, B. Wu, Y. Shao, J. Electroanal. Chem. 91. W. Nernst, Z. Phys. Chem. (Frankfurt). 2, 613, (1888).
553, 43, (2003). 92. W. Nernst, Z. Phys. Chem. (Frankfurt). 4, 129, (1889).
47. M. V. Manzanares, Rev. Soc. Quim. Mex. 47, 66, (2003). 93. W. Nernst, E. H. Riesenfeld, Ann. Phys. 8, 600, (1902).
48. Y. Yatziv, I. Turyan, D. Mandler, J. Am. Chem. Soc. 124, 5618, (2002). 94. P. Bothorel, C. Lussan, C. R. l’Academie. Sci. 266, 2492, (1968).
49. Y. Kitatsuji, Z. Yoshida, H. Kudo, S. Kihara, J. Electroanal. Chem. 520, 95. W. D. Harkins, E. C. Humphery, J. Am. Chem. Soc. 38, 228, (1916).
133, (2002). 96. G. C. H. Ehrensvard, D. F. Cheesman, Svensk Kem. Tid. 53, 126,
50. J. Koryta, W. Ruth, P. Vanysek, A. Hofmanova, Anal. Lett. 15, 1685, (1941).
(1982). 97. E. K. Fischer, W. D. Harkins, J. Phys. Chem. 36, 98, (1932).
51. G. Du, J. Koryta, W. Ruth, P. Vanysek, J. Electroanal. Chem. 159, 413, 98. T. R. Bolam, J. Crowe, J. Phys. Chem. 35, 1448, (1931).
(1983). 99. R. P. Bell, J. Phys. Chem. 32, 882, (1928).
52. V. Marecek, H. Janchenova, M. Brezina, M. Betti, Anal. Chim. Acta. 100. P. L. du Nouy, C. R. l’Academie. Sci. 180, 1579, (1925).
244, 15, (1991). 101. J. H. Mathews, A. J. Stamm, J. Am. Chem. Soc. 46, 1071, (1924).
53. Z. Pang, D. Guo, X. Teng, E. Wang, Chin. J. Pharm. Anal. 5, 209, 102. W. D. Harkins, E. C. Humphery, J. Am. Chem. Soc. 38, 236, (1916).
(1985). 103. J. Koryta, P. Vanysek, M. Brezina, J. Electroanal. Chem. 67, 263,
54. J. Koryta, G. Du, W. Ruth, P. Vanysek, Faraday Discuss., 209, (1984). (1976).
55. A. Sabela, J. Koryta, O. Valent, J. Electroanal. Chem. 204, 267, (1986). 104. J. Koryta, P. Vanysek, M. Brezina, J. Electroanal. Chem. 75, 211,
56. J. Koryta, Wiad. Chem. 44, 579, (1990). (1977).
57. P. Vanysek, W. Ruth, J. Koryta, J. Electroanal. Chem. 148, 117, (1983). 105. P. Vanysek, J. Electroanal. Chem. 121, 149, (1981).
58. I. Koryta, N. Kozlov Iu, A. Gofmanova, V. Khalil, P. Vanysek, Antibiotiki. 106. J. Koryta, M. Brezina, A. Hofmanova, D. Homolka, H. Le Quoc, W.
28, 810, (1983). Khalil, V. Marecek, Z. Samec, S. K. Sen, et al., Bioelectrochem.
59. H. Ohde, K. Maeda, O. Shirai, Y. Yoshida, S. Kihara, J. Electroanal. Bioenerget. 7, 61, (1980).
Chem. 438, 139, (1997). 107. P. Vanysek, M. Behrendt, J. Electroanal. Chem. 130, 287, (1981).
60. E. Makrlik, P. Vanura, J. Halova, Pol. J. Chem. 81, 1531, (2007). 108. D. Homolka, L. Q. Hung, A. Hofmanova, M. W. Khalil, J. Koryta, V.
61. E. Makrlik, P. Vanura, Acta Chim. Slov. 54, 375, (2007). Marecek, Z. Samec, S. K. Sen, P. Vanysek, J. Weber, Anal. Chem. 52,
62. E. Makrlik, P. Vanura, J. Radioanal. Nucl. Chem. 268, 641, (2006). 1606, (1980).
63. E. Makrlik, P. Vanura, J. Radioanal. Nucl. Chem. 268, 155, (2006). 109. Z. Samec, V. Marecek, D. Homolka, J. Electroanal. Chem. 126, 121,
64. L. Basaez, P. Vanysek, J. Pharm. Biomed. Anal. 19, 183, (1999). (1981).
65. D. P. Tieleman, H. J. C. Berendsen, M. S. P. Sansom, Biophys. J. 80, 331, 110. T. Wandlowski, S. Racinsky, V. Marecek, Z. Samec, J. Electroanal.
(2001). Chem. 227, 281, (1987).
66. A. Illanes, L. Wilson, O. Corrotea, L. Tavernini, F. Zamorano, C. Aguirre, 111. T. Wandlowski, V. Marecek, Z. Samec, J. Electroanal. Chem. 242, 277,
J. Mol. Catal. B Enzym. 47, 72, (2007). (1988).
67. C. Aguirre, P. Opazo, M. Venegas, R. Riveros, A. Illanes, Process 112. P. Vanysek, I. C. Hernandez, J. Xu, Microchem. J. 41, 327, (1990).
Biochem. (Amsterdam). 41, 1924, (2006). 113. J. D. Reid, O. R. Melroy, R. P. Buck, J. Electroanal. Chem. 147, 71,
68. T. Kondo, T. Kakiuchi, M. Senda, Biochim. Biophys. Acta. 1, (1992). (1983).
69. L. Murtomaki, J. A. Manzanares, S. Mafe, K. Kontturi, Surf. Sci. Ser. 95, 114. V. Marecek, Z. Samec, Anal. Chim. Acta. 151, 265, (1983).
533, (2001). 115. V. Marecek, Z. Samec, J. Electroanal. Chem. 149, 185, (1983).
70. Z. Samec, A. Trojanek, P. Krtil, Faraday Discuss. 129, 301, (2005). 116. T. Kakiuchi, Y. Takasu, Anal. Chem. 66, 1853, (1994).
71. J. Koryta, L. Q. Hung, A. Hofmanova, Stud. Biophys. 90, 25, (1982). 117. T. Kakiuchi, Y. Takasu, J. Electroanal. Chem. 365, 293, (1994).
72. Z. Samec, A. Trojanek, H. H. Girault, Electrochem. Commun. 5, 98, 118. T. Kakiuchi, Y. Takasu, J. Electroanal. Chem. 381, 5, (1995).
(2003). 119. H. Nagatani, S. Suzuki, D. J. Fermin, H. H. Girault, K. Nakatani, Anal.
73. H. H. J. Girault, D. J. Schiffrin, J. Electroanal. Chem. 179, 277, (1984). Bioanal. Chem. 386, 633, (2006).
74. T. Kakiuchi, M. Kotani, J. Noguchi, M. Nakanishi, M. Senda, J. Colloid 120. J. A. Campbell, H. H. Girault, J. Electroanal. Chem. 266, 465, (1989).
Interface Sci. 149, 279, (1992). 121. A. A. Stewart, G. Taylor, H. H. Girault, J. McAleer, J. Electroanal.
Chem. 296, 491, (1990).

1461
J. Chil. Chem. Soc., 53, Nº 2 (2008)

122. A. A. Stewart, Y. Shao, C. M. Pereira, H. H. Girault, J. Electroanal. 166. A. Maelkiae, P. Liljeroth, A.-K. Kontturi, K. Kontturi, J. Phys. Chem. B.
Chem. 305, 135, (1991). 105, 10884, (2001).
123. Y. Shao, H. H. Girault, J. Electroanal. Chem. 334, 203, (1992). 167. S. A. Dassie, A. M. Baruzzi, J. Electroanal. Chem. 492, 94, (2000).
124. T. Ohkouchi, T. Kakutani, T. Osakai, M. Senda, Anal. Sci. 7, 371, 168. M. J. Tupy, H. W. Blanch, C. J. Radke, Ind. Eng. Chem. Res. 37, 3159,
(1991). (1998).
125. V. Sladkov, V. Guillou, S. Peulon, M. L’Her, J. Electroanal. Chem. 573, 169. G. W. Stevens, J. M. Perera, F. Grieser, Curr. Opin. Colloid Interface.
129, (2004). Sci. 2, 629, (1997).
126. R. M. Lahtinen, D. J. Fermin, H. Jensen, K. Kontturi, H. H. Girault, 170. J. Plesek, B. Gruner, S. Hermanek, J. Baca, V. Marecek, J. Janchenova,
Electrochem. Commun. 2, 230, (2000). A. Lhotsky, K. Holub, P. Selucky, J. Rais, I. Cisarova, J. Caslavsky,
127. D. J. Fermin, R. Lahtinen, Surf. Sci. Ser. 95, 179, (2001). Polyhedron. 21, 975, (2002).
128. H. H. J. Girault, D. J. Schiffrin, J. Electroanal. Chem. 244, 15, (1988). 171. C. Forssten, J. Strutwolf, D. E. Williams, Electrochem. Commun. 3, 619,
129. G. Geblewicz, D. J. Schiffrin, J. Electroanal. Chem. 244, 27, (1988). (2001).
130. M. L’Her, V. Sladkov, Electron transfer at the liquid-liquid interface, in 172. R. Gulaboski, M. N. D. S. Cordeiro, N. Milhazes, J. Garrido, F. Borges,
Trends in Molecular Electrochemistry, A. J. L. Pombeiro, C. Amatore, M. Jorge, C. M. Pereira, I. Bogeski, A. H. Morales, B. Naumoski, A. F.
Editors. CRC Press, Boca Raton, 2004, p. 503. Silva, Anal. Biochem. 361, 236, (2007).
131. T. Solomon, A. J. Bard, J. Phys. Chem. 99, 17487, (1995). 173. L. Basaez, I. Peric, C. Aguirre, P. Vanysek, Bol. Soc. Chil. Quim. 46, 203,
132. R. A. Marcus, J. Phys. Chem. 94, 7742, (1990). (2001).
133. R. A. Marcus, J. Phys. Chem. 94, 4152, (1990). 174. H. Alemu, Pure Appl. Chem. 76, 697, (2004).
134. Y. Shao, M. V. Mirkin, J. F. Rusling, J. Phys. Chem. B. 101, 3202, 175. F. Reymond, H. J. Lee, J. S. Rossier, L. Tomaszewski, R. Ferrigno, C. M.
(1997). Pereira, H. H. Girault, Chimia. 53, 103, (1999).
135. F. G. Chevallier, T. J. Davies, O. V. Klymenko, L. Jiang, T. G. J. Jones, 176. J. Koryta, Selective Electr. Rev. 13, 133, (1991).
R. G. Compton, J. Electroanal. Chem. 580, 265, (2005). 177. B. Liu, M. V. Mirkin, Electroanalysis. 12, 1433, (2000).
136. B. Su, N. Eugster, H. H. Girault, J. Electroanal. Chem. 577, 187, 178. Z. Samec, E. Samcova, H. H. Girault, Talanta. 63, 21, (2004).
(2005). 179. B. J. Seddon, Y. Shao, J. Fost, H. H. Girault, Electrochim. Acta. 39, 783,
137. J. Zhang, C. J. Slevin, L. Murtomaeki, K. Kontturi, D. E. Williams, P. R. (1994).
Unwin, Langmuir. 17, 821, (2001). 180. D. Zhan, X. Li, W. Zhan, F.-R. F. Fan, A. J. Bard, Anal. Chem. 79, 5225,
138. R. R. Naujok, D. A. Higgins, D. G. Hanken, R. M. Corn, J. Chem. Soc., (2007).
Faraday Trans. 91, 1411, (1995). 181. D. A. Walsh, J. L. Fernandez, J. Mauzeroll, A. J. Bard, Anal. Chem. 77,
139. H. H. Girault, Charge transfer across liquid-liquid interfaces, in Modern 5182, (2005).
Aspects of Electrochemistry, J. O. M. Bockris, B. E. Conway, R. E. 182. P. Sun, Z. Zhang, Z. Gao, Y. Shao, Angew. Chem. Int. Ed. 41, 3445,
White, Editors. Plenum Publishing Corp., New York, 1993, p. 1. (2002).
140. M. Vignali, R. Edwards, V. J. Cunnane, J. Electroanal. Chem. 592, 37, 183. Q. Qian, G. S. Wilson, K. Bowman-James, Electroanalysis. 16, 1343,
(2006). (2004).
141. C. Johans, R. Lahtinen, K. Kontturi, D. J. Schiffrin, J. Electroanal. 184. Q. Qian, G. S. Wilson, K. Bowman-James, H. H. Girault, Anal. Chem.
Chem. 488, 99, (2000). 73, 497, (2001).
142. M. Platt, R. A. W. Dryfe, J. Electroanal. Chem. 599, 323, (2007). 185. R. Zazpe, C. Hibert, J. O’Brien, Y. H. Lanyon, D. W. M. Arrigan, Lab
143. Y. Luan, M. An, G. Lu, Appl. Surf. Sci. 253, 459, (2006). Chip. 7, 1732, (2007).
144. M. Vignali, R. A. H. Edwards, M. Serantoni, V. J. Cunnane, J. Electroanal. 186. F. Quentel, C. Elleouet, V. Mirceski, V. Agmo Hernandez, M. L’Her, M.
Chem. 591, 59, (2006). Lovric, S. Komorsky-Lovric, F. Scholz, J. Electroanal. Chem. 611, 192,
145. V. Mirceski, R. Gulaboski, J. Phys. Chem. B. 110, 2812, (2006). (2007).
146. V. Mirceski, Electrochem. Commun. 8, 123, (2006). 187. J. Langmaier, Z. Samec, Electrochem. Commun. 9, 2633, (2007).
147. I. Turyan, M. Etienne, D. Mandler, W. Schuhmann, Electroanalysis. 17, 188. R. Ishimatsu, N. Nishi, T. Kakiuchi, Langmuir. 23, 7608, (2007).
538, (2005). 189. R. Ishimatsu, N. Nishi, T. Kakiuchi, Chem. Lett. 36, 1166, (2007).
148. M. Platt, R. A. W. Dryfe, Phys. Chem. Chem. Phys. 7, 1807, (2005). 190. T. Kakiuchi, Anal. Chem. 79, 6442, (2007).
149. E. Tada, Y. Oishi, H. Kaneko, Electrochem. Solid State Lett. 8, C26, 191. V. A. Hernandez, F. Scholz, Electrochem. Commun. 8, 967, (2006).
(2005). 192. K. Tanaka, N. Nishi, T. Kakiuchi, Anal. Sci. 20, 1553, (2004).
150. M. Platt, R. A. W. Dryfe, E. P. L. Roberts, Electrochim. Acta. 49, 3937, 193. W. Schmickler, Interfacial Electrochemistry. Oxford University Press.
(2004). New York. 1996, 304 pp.
151. C. R. C. Johans, Novel routes to metal nanoparticles: electrodeposition 194. K. L. Kott, D. A. Higgins, R. J. McMahon, R. M. Corn, J. Am. Chem.
and reactions at liquid/liquid interfaces, in Department of Chemical Soc. 115, 5342, (1993).
Technology. 2003, Helsiki University of Technology: Espoo, Finland. p. 195. D. A. Higgins, R. M. Corn, J. Phys. Chem. 97, 489, (1993).
56 pp. 196. D. A. Higgins, R. R. Naujok, R. M. Corn, Chem. Phys. Lett. 213, 485,
152. A. Eftekhari, Appl. Surf. Sci. 227, 331, (2004). (1993).
153. E. Tada, H. Kaneko, Chem. Lett., 1306, (2000). 197. H. Nagatani, A. Piron, P.-F. Brevet, D. J. Fermin, H. H. Girault, Langmuir.
154. M. Guainazzi, G. Silvestri, G. Serravalle, J. Chem. Soc., Chem. Commun., 18, 6647, (2002).
200, (1975). 198. J. G. Frey, Second harmonic generation at liquid/liquid interfaces, in
155. P. Vanysek, L. A. Delia, Electroanalysis. 18, 371, (2006). Interfacial Nanochemistry, H. Watari, N. Teramae, T. Sawada, Editors.
156. P. Vanysek, Proceedings of the IEEE International Frequency Control Springer, New York, 2005, p. 1.
Symposium. 51, 49, (1997). 199. J. C. Conboy, G. L. Richmond, Electrochim. Acta. 40, 2881, (1995).
157. T. Solomon, A. J. Bard, Anal. Chem. 67, 2787, (1995). 200. J. C. Conboy, G. L. Richmond, J. Phys. Chem. B. 101, 983, (1997).
158. T. Solomon, A. J. Bard, Bull. Chem. Soc. Ethiopia. 11, 55, (1997). 201. H. H. Girault, M. Hojeij, B. Su, Abstracts of Papers, 233rd ACS National
159. C. Wei, A. J. Bard, M. V. Mirkin, J. Phys. Chem. 99, 16033, (1995). Meeting, Chicago, IL, United States, March 25-29, 2007. PHYS,
160. M. Tsionsky, A. J. Bard, M. V. Mirkin, J. Phys. Chem. 100, 17881, (2007).
(1996). 202. J. C. Conboy, G. L. Richmond, Total internal reflection second harmonic
161. Y. Selzer, D. Mandler, J. Electroanal. Chem. 409, 15, (1996). generation from the interface between two immiscible electrolyte
162. P. Vanysek, Anal. Chem. 62, 827A, (1990). solutions. 1995, Dep. Chem., Oregon Univ., Eugene, OR, USA. p. 15
163. F. Reymond, D. Fermin, H. J. Lee, H. H. Girault, Electrochim. Acta. 45, pp.
2647, (2000). 203. R. M. Corn, R. N. Naujok, H. J. Paul, Book of Abstracts, 210th ACS
164. R. B. Gennis, Biomembranes: Molecular Structure and Function. National Meeting, Chicago, IL, August 20-24. COLL, (1995).
Springer. New York. 1988, 533 pp. 204. S. Ishizaka, N. Kitamura, Bull. Chem. Soc. Japan. 74, 1983, (2001).
165. T. Spataru, N. Spataru, N. Bonciocat, C. Luca, Bioelectrochemistry. 62, 205. M. J. Tupy, Protein adsorption dynamics at the oil/water interface as
67, (2004). studied by total internal reflection fluorescence spectroscopy. 1998,
University of California at Berkeley: Berkeley. p. 268 pp.

1462
J. Chil. Chem. Soc., 53, Nº 2 (2008)

206. J. M. Perera, G. W. Stevens, F. Grieser, Colloids Surf., A. 95, 185, 239. S. Frank, W. Schmickler, J. Electroanal. Chem. 483, 18, (2000).
(1995). 240. I. Benjamin, Science. 261, 1558, (1993).
207. J. C. Conboy, J. L. Daschbach, G. L. Richmond, J. Phys. Chem., 9688, 241. I. Benjamin, Chem. Rev. 96, 1449, (1996).
(1994). 242. L. Benjamin, J. Chem. Phys. 97, 1432, (1992).
208. D. S. Walker, M. G. Brown, C. L. McFearin, G. L. Richmond, J. Phys. 243. D. Michael, I. Benjamin, J. Electroanal. Chem. 450, 335, (1998).
Chem. B. 108, 2111, (2004). 244. P. A. Fernandes, M. N. D. S. Cordeiro, J. A. N. F. Gomes, J. Phys. Chem.
209. H. Yui, Y. Ikezoe, T. Sawada, Anal. Sci. 20, 1501, (2004). B. 103, 8930, (1999).
210. J. Strutwolf, A. L. Barker, M. Gonsalves, D. J. Caruana, P. R. Unwin, D. 245. M. R. Philpott, T.-T. Lin, J. N. Glosli. Molecular Dynamics Study of
E. Williams, J. R. P. Webster, J. Electroanal. Chem. 483, 163, (2000). Neutral and Charged Liquid-Liquid Interfaces. in Modeling of Processes
211. J. R. P. Webster, Abstracts of Papers, 234th ACS National Meeting, at Electrochemical Interfaces and in Electrochemical Systems. 1999.
Boston, MA, United States, August 19-23, 2007. PHYS, (2007). Seattle: The Electrochemical Society.
212. A. Zarbakhsh, A. Querol, J. Bowers, M. Yaseen, J. R. Lu, J. R. P. Webster, 246. P. Tarazona, R. Checa, E. Chacon, Phys. Rev. Lett. 99, 196101/1,
Langmuir. 21, 11704, (2005). (2007).
213. A. Zarbakhsh, A. Querol, J. Bowers, J. R. P. Webster, Faraday Discuss. 247. J. Chowdhary, B. M. Ladanyi, Abstracts of Papers, 232nd ACS National
129, 155, (2005). Meeting, San Francisco, CA, United States, Sept. 10-14, 2006. COLL,
214. G. Luo, S. Malkova, S. V. Pingali, D. G. Schultz, B. Lin, M. Meron, I. (2006).
Benjamin, P. Vanysek, M. L. Schlossman, J. Phys. Chem. B. 110, 4527, 248. D. M. Mitrinovic, A. M. Tikhonov, M. Li, Z. Huang, M. L. Schlossman,
(2006). Phys. Rev. Lett. 85, 582, (2000).
215. G. Luo, S. Malkova, S. V. Pingali, D. G. Schultz, B. Lin, M. Meron, 249. A. Zarbakhsh, J. Bowers, J. R. P. Webster, Langmuir. 21, 11596, (2005).
T. J. Graber, J. Gebhardt, P. Vanysek, M. L. Schlossman, Electrochem. 250. M. L. Schlossman, Curr. Opin. Colloid Interface. Sci. 7, 235, (2002).
Commun. 7, 627, (2005). 251. C. W. Monroe, M. Urbakh, A. A. Kornyshev, J. Phys.: Condens. Matter.
216. G. Luo, S. Malkova, S. V. Pingali, D. G. Schultz, B. Lin, M. Meron, T. 19, 375113/1, (2007).
J. Graber, J. Gebhardt, P. Vanysek, M. L. Schlossman, Faraday Discuss. 252. R. Chertcoff, A. Calvo, I. Paterson, M. Rosen, J. P. Hulin, J. Colloid
129, 23, (2005). Interface Sci. 154, 194, (1992).
217. G. Luo, S. Malkova, a. Yoon, D. G. Schultz, B. Lin, M. Meron, I. 253. P. Joos, R. V. Bogaert, J. Colloid Interface Sci. 56, 213, (1976).
Benjamin, P. Vanysek, M. L. Schlossman, Science. 311, 216, (2006). 254. M. Dupeyrat, J. Michel, Experientia Suppl. 18, 269, (1971).
218. G. Luo, S. Malkova, J. Yoon, D. Schultz, B. Lin, I. Benjamin, P. Vanysek, 255. M. Dupeyrat, E. Nakache, C. R. Acad. Sci. (Paris). 277, 599, (1973).
M. L. Schlossman, Abstracts of Papers, 233rd ACS National Meeting, 256. E. Nakache, M. Dupeyrat, M. Vignes-Adler, Faraday Discuss. Chem.
Chicago, IL, United States, March 25-29, 2007. PHYS, (2007). Soc. 77, 189, (1984).
219. G. Luo, S. Malkova, J. Yoon, D. G. Schultz, B. Lin, M. Meron, I. 257. P. Van Remoortere, P. Joos, J. Colloid Interface Sci. 160, 397, (1993).
Benjamin, P. Vanysek, M. L. Schlossman, J. Electroanal. Chem. 593, 258. A. A. Kornyshev, A. M. Kuznetsov, M. Urbakh, J. Chem. Phys. 117,
142, (2006). 6766, (2002).
220. T. Kakiuchi, T. Takaishi, Proceedings - Electrochemical Society. 97-19, 259. A. L. Barker, P. R. Unwin, J. Phys. Chem. B. 105, 12019, (2001).
538, (1997). 260. Gouy, C. R. l’Academie. Sci. 149, 654, (1910).
221. A. A. Stewart, J. A. Campbell, H. H. Girault, M. Eddowes, Ber. Bunsen- 261. D. L. Chapman, Phil. Mag. Ser. 6. 25, 475, (1913).
Ges. Phys. Chem. 94, 83, (1990). 262. S. Levine, C. W. Outhwaite, L. B. Bhuiyan, J. Electroanal. Chem. 123,
222. X. Lu, H. Zhang, L. Hu, C. Zhao, L. Zhang, X. Liu, Electrochem. 105, (1981).
Commun. 8, 1027, (2006). 263. C. M. Pereira, A. Martins, M. Rocha, C. J. Silva, F. Silva, J. Chem. Soc.,
223. X. Lu, L. Hu, X. Wang, Electroanalysis. 17, 953, (2005). Faraday Trans. 90, 143, (1994).
224. M. V. Mirkin, M. Tsionsky, Charge-transfer at the liquid/liquid interface, 264. Y. Cheng, V. J. Cunnane, D. J. Schiffrin, L. Murtomaki, K. Kontturi, J.
in Scanning Electrochemical Microscopy, A. J. Bard, M. V. Mirkin, Chem. Soc., Faraday Trans. 87, 1665, (1991).
Editors. CRC Press, Boca Raton, 2001, p. 299. 265. Y. Cheng, V. J. Cunnane, D. J. Schiffrin, L. Mutomaki, K. Kontturi, J.
225. J. Zhang, P. R. Unwin, Phys. Chem. Chem. Phys. 4, 3820, (2002). Chem. Soc., Faraday Trans. 87, 107, (1991).
226. A. L. Barker, P. R. Unwin, J. Zhang, Electrochem. Commun. 3, 372, 266. C. M. Pereira, W. Schmickler, F. Silva, M. J. Sousa, J. Electroanal.
(2001). Chem. 436, 9, (1997).
227. B. Quinn, K. Kontturi, J. Electroanal. Chem. 483, 124, (2000). 267. O. Pecina, J. P. Badiali, Phys. Rev. E. Sta. Phys. Plasma Fluid. Rel. 58,
228. J. Zhang, A. L. Barker, P. R. Unwin, J. Electroanal. Chem. 483, 95, 6041, (1998).
(2000). 268. H. H. Girault, D. J. Schiffrin, Electrochemistry of liquid-liquid interfaces,
229. A. L. Barker, P. R. Unwin, J. Phys. Chem. B. 104, 2330, (2000). in Electroanalytical Chemistry, A. J. Bard, Editor. Marcel Dekker, New
230. T. Kakiuchi, M. Senda, Bull. Chem. Soc. Japan. 56, 1322, (1983). York, 1989, p. 1.
231. C. Gavach, P. Seta, B. D’Epenoux, J. Electroanal. Chem. 83, 225, 269. O. Stern, Z. Elektrochem. 30, 508, (1924).
(1977). 270. P. Vanysek, TrAC, Trends Anal. Chem. 12, 357, (1993).
232. M. Gros, S. Gromb, C. Gavach, J. Electroanal. Chem. 89, 29, (1978). 271. A. Volkov, D. W. Deamer, eds. Liquid-Liquid Interfaces. Theory and
233. Z. Samec, V. Marecek, D. Homolka, J. Electroanal. Chem. 187, 31, Methods. 1st ed. CRC, Boca Raton, 1996. 448 pp.
(1985). 272. A. G. Volkov, V. S. Markin, Electrochemical double layers: liquid-liquid
234. H. H. Girault, D. J. Schiffrin, J. Electroanal. Chem. 150, 43, (1983). interfaces, in Encyclopedia of Electrochemistry. Wiley-VCH, New York,
235. H. H. Girault, Electrochim. Acta. 32, 383, (1987). 2003, p. 162.
236. H. H. J. Girault, D. J. Schiffrin, J. Electroanal. Chem. 170, 127, (1984). 273. A. G. Volkov, V. S. Markin, Electric properties of oil/water interfaces,
237. C. M. Pereira, W. Schmickler, A. F. Silva, M. J. Sousa, Chem. Phys. Lett. in Emulsions: Structure, stability and interactions, 4, D. Petsev, Editor.
268, 13, (1997). Academic Press, New York, 2004, p. 91.
238. D. J. Henderson, W. Schmickler, J. Chem. Soc., Faraday Trans. 92,
3839, (1996).

1463

Das könnte Ihnen auch gefallen